首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
We have developed a boronate affinity immunoassay system using m-aminophenylboronic acid (mAPB) coupling to bacterial magnetic particles (BMPs). Homobifunctional crosslinker, Bis-(succcimidyl)suberate (BS3), was employed for preparation of mAPB-BMPs conjugates (mAPB-BMPs). Quantities of HbA1c on mAPB-BMPs were evaluated based on luminescence from alkaline phosphatase-conjugated anti-Hb antibody (ALP–antibody) binding to HbA1c on the BMP surface. The binding of HbA1c to mAPB-BMPs occurred gradually and was almost completed within 10 mm. The coupling reaction is enhanced due to static electric interaction between the positive charges on HbA1c and negative charges on BMPs. The amount of HbA1c binding to mAPB-BMPs increased with increasing sodium chloride concentrations in the range of 0–100 mM. However, the amount of Hb binding to mAPB-BMPs also increased in high concentration of sodium chloride. The Hb binding to mAPB-BMPs was detached from mAPB-BMPs when Hb–mAPB-BMPs were washed with low salt buffer. This indicates that Hb is nonspecifically adsorbed onto the surface of mAPB-BMPs in high concentration of sodium chloride. These results suggest that selective separation of HbA1c using mAPB-BMPs can be achieved with these conditions. A dose–response curve was obtained between luminescence intensity and HbA1c concentration using a fully automated boronate affinity immunoassay. A linear relationship between luminescence intensity and HbA1c concentration was obtained in the range of 10–104 ng/ml.  相似文献   

2.
R. J. Reid  L. D. Field  M. G. Pitman 《Planta》1985,166(3):341-347
31P-Nuclear magnetic resonance spectroscopy was used to measure the cytoplasmic pH (pHc) in barley (Hordeum vulgare L.) root tips. As the external pH was raised from 4–10, pHc was found to increase from 7.44 to 7.75. The sensitivity of pHc to changes in external pH decreased with increasing external pH. Metabolic inhibition by sodium azide caused pHc to fall by 0.3 units. Addition of 10 mM butyrate resulted in a gradual decline in pHc, by approx. 0.3 units over 90 min. At a concentration of 1 mM, butyrate had no effect on pHc even after 2 h. Fusicoccin caused pHc to rise by 0.1–0.2 units. In maize (Zea mays L.) root tips, pHc was shown to have a similar sensitivity to fusicoccin. The results are discussed in relation to the regulation of pHc and the possible role of pHc in determining transmembrane electrical potential differences.Abbreviations and symbols FC Fusicoccin - NMR nuclear magnetic resonance - p.d. membrane electrical potential difference - pHc cytoplasmic pH - P1 inorganic phosphate - chemical shift  相似文献   

3.
Flash-induced kinetics of the membrane potential increase related to electron transfer within the cytochrome (cyt) b/c1 complex (Phase III) and that of cyt c1+c2 reduction have been measured as a function of myxothiazol concentration in isolated chromatophores and whole cells of Rhodobacter sphaeroides. Upon addition of nonsaturating concentrations of myxothiazol, kinetics of Phase III display two phases, Phase IIIa and Phase IIIb. The amplitude of Phase IIIa, completed in about 10 ms, is proportional to the fraction of non-inhibited cyt b/c1 complexes, while its half-time is independent of the myxothiazol concentration. A fast cyt c1+c2 reduction phase is correlated to Phase IIIa. These experiments demonstrate that, in a range of time of several ms, diffusion of cyt c2 is restricted to domains formed by a supercomplex including two reaction centers (RCs) and a single cyt b/c1 complex, as proposed by Joliot et al. (Biochim Biophys Acta 975: 336–345, 1989). Phase IIIb, completed in about 100 ms, shows that positive charges or inhibitor molecules are exchanged between supercomplexes in this range of time. These exchanges occur within domains including 2 to 3 supercomplexes, i.e. in membrane domains smaller than a single chromatophore. These conclusions apply to both isolated chromatophores and whole cells.Abbreviations cyt cytochrome - MOPS 3-(N-morpholino)propane sulfonic acid - PMS phenazine methosulfate - P primary donor - Rb. Rhodobacter - RC reaction center  相似文献   

4.
Fructosyl peptide oxidase is a flavoenzyme that catalyzes the oxidative deglycation of N-(1-deoxyfructosyl)-Val-His, a model compound of hemoglobin (Hb)A1C. To develop an enzymatic method for the measurement of HbA1C, we screened for a proper protease using N-(1-deoxyfructosyl)-hexapeptide as a substrate. Several proteases, including Neutral protease from Bacillus polymyxa, were found to release N-(1-deoxyfructosyl)-Val-His efficiently, however no protease was found to release N-(1-deoxyfructosyl)-Val. Neutral protease also digested HbA1C to release N-(1-deoxyfructosyl)-Val-His, and then the fructosyl peptide was detected using fructosyl peptide oxidase. The linear relationship was observed between the concentration of HbA1C and the absorbancy of fructosyl peptide oxidase reaction, hence this new method is a practical means for measuring HbA1C.  相似文献   

5.
We have separated and quantified two new minor hemoglobins named HbA1d3a and HbA1d3b. The level of HbA1d3a was significantly higher in uremic than in non-uremic patients (3.00 ± 0.50% vs. 1.28 ± 0.26% of total hemoglobin). It correlated well with carbamylated hemoglobin (r=0.80, n=81, p<0.002) and with plasma urea concentration (r=0.78, n=81, p<0.002). These data and the electrospray ionization mass spectrometric analysis provide strong evidence that HbA1d3a is an α-chain modified by carbamylation. The HbA1d3b level in the diabetic patients was found to be 1.6-fold that in non-diabetic subjects (3.00 ± 0.49 vs. 1.90 ± 0.33). This was attributed to HbA1d3 modified by glycation. Indeed HbA1d3b correlated significantly with HbA1c (r=0.71, p<0.002) and with serum glucose level (r=0.62, p<0.002). These two new minor hemoglobins may serve as complements for the objective assessment of averagd long-term uremia and glycemia in uremic and diabetic patients.  相似文献   

6.
Naphthylphthalamic acid (NPA), an inhibitor of polar auxin transport, binds with high affinity to membrane preparations from callus and cell suspension cultures derived from Nicotiana tabacum (K d approx. 2·10–9 M). The concentration of membrane-bound binding sites is higher in cell suspension than in callus cultures. The binding of NPA to these sites seems to be a simple process, in contrast to the binding of the synthetic auxin naphthylacetic acid (1-NAA) to membrane preparations from callus cultures, which is more complex (A.C. Maan et al., 1983, Planta 158, 10–15). Naphthylacetic acid, a number of structurally related compounds and the auxin-transport inhibitor triiodobenzoic acid were all able to compete with NPA for the same binding site with K d values ranging from 10–6 to 10–4 M. On the other hand, NPA was not able to displace detectable amounts of NAA from the NAA-binding site. A possible explantation is the existence of two different membrane-bound binding sites, one exclusively for auxins and one for NPA as well as auxins, that differ in concentration. The NPA-binding site is probably an auxin carrier.Abbreviations 1-NAA 1-Naphthylacetic acid - 2-NAA 2-Naphthylacetic acid - NPA N-1-Naphthylphthalamic acid  相似文献   

7.
Glycated hemoglobin (HbA1c) is a ‘gold standard’ biomarker for assessing the glycemic index of an individual. HbA1c is formed due to nonenzymatic glycosylation at N-terminal valine residue of the β-globin chain. Cation exchange based high performance liquid chromatography (CE–HPLC) is mostly used to quantify HbA1c in blood sample. A few genetic variants of hemoglobin and post-translationally modified variants of hemoglobin interfere with CE–HPLC-based quantification, resulting in its false positive estimation. Using mass spectrometry, we analyzed a blood sample with abnormally high HbA1c (52.1%) in the CE–HPLC method. The observed HbA1c did not corroborate the blood glucose level of the patient. A mass spectrometry based bottom up proteomics approach, intact globin chain mass analysis, and chemical modification of the proteolytic peptides identified the presence of Hb Beckman, a genetic variant of hemoglobin, in the experimental sample. A similar surface area to charge ratio between HbA1c and Hb Beckman might have resulted in the coelution of the variant with HbA1c in CE–HPLC. Therefore, in the screening of diabetes mellitus through the estimation of HbA1c, it is important to look for genetic variants of hemoglobin in samples that show abnormally high glycemic index, and HbA1c must be estimated using an alternative method.  相似文献   

8.
Mitochondrial cytochromec (horse), which is a very efficient electron donor to bacterial photosynthetic reaction centersin vitro, binds to the reaction center ofRhodospirillum rubrum with an approximate dissociation constant of 0.3–0.5 µM at pH 8.2 and low ionic strength. The binding site for the reaction center is on the frontside of cytochromec which is the side with the exposed heme edge, as revealed by differential chemical acetylation of lysines of free and reaction-center-bound cytochromec. In contrast, bacterial cytochromec 2 was found previously to bind to the detergent-solubilized reaction center through its backside, i.e., the side opposite to the heme cleft [Rieder, R., Wiemken, V., Bachofen, R., and Bosshard, H. R. (1985).Biochem. Biophys. Res. Commun. 128, 120–126]. Binding of mitochondrial cytochromec but not of mitochondrial cytochromec 2 is strongly inhibited by low concentrations of poly-l-lysine. The results are difficult to reconcile with the existence of an electron transfer site on the backside of cytochromec 2.  相似文献   

9.
Cytochrome a 1 c 1 was highly purified from Nitrobacter agilis. The cytochrome contained heme a and heme c of equimolar amount, and its reduced form showed absorption peaks at 587, 550, 521, 434 and 416 nm. Molecular weight per heme a of the cytochrome was estimated to be approx. 100,000–130,000 from the amino acid composition. A similar value was obtained by determining the protein content per heme a. The cytochrome molecule was composed of three subunits with molecular weights of 55,000, 29,000 and 19,000, respectively. The 29 kd subunit had heme c.Hemes a and c of cytochrome a 1 c 1 were reduced on addition of nitrite, and the reduced cytochrome was hardly autoxidizable. Exogenously added horse heart cytochrome c was reduced by nitrite in the presence of cytochrome a 1 c 1; K m values of cytochrome a 1 c 1 for nitrite and N. agilis cytochrome c were 0.5 mM and and 6 M, respectively. V max was 1.7 mol ferricytochrome c reduced/min·mol of cytochrome a 1 c 1 The pH optimum of the reaction was about 8. The nitrite-cytochrome c reduction catalyzed by cytochrome a 1 c 1 was 61% and 88% inhibited by 44M azide and cyanide, respectively. In the presence of 4.4 mM nitrate, the reaction was 89% inhibited. The nitrite-cytochrome c reduction catalysed by cytochrome a 1 c 1 was 2.5-fold stimulated by 4.5 mM manganous chloride. An activating factor which was present in the crude enzyme preparation stimulated the reaction by 2.8-fold, and presence of both the factor and manganous ion activated the reaction by 7-fold.Cytochrome a 1 c 1 showed also cytochrome c-nitrate reductase activity. The pH optimum of the reaction was about 6. The nitrate reductase activity was also stimulated by manganous ions and the activating factor.  相似文献   

10.
Turnover of the ubiquinol oxidizing site of the UQH2:cyt c2 oxidoreductase (b/c 1 complex) ofRps. sphaeroides can be assayed by measuring the rate of reduction of cytb 561 in the presence of antimycin (AA). Oxidation of ubiquinol is a second-order process, with a value ofk 2 of about 3 × 105 M–1. The reaction shows saturation at high quinol concentrations, with an apparentK m of about 6–8 mM (with respect to the concentration of quinol in the membrane). When the quinone pool is oxidized before illumination, reduction of the complex shows a substantial lag (about 1 ms) after a flash, indicating that the quinol produced as a result of the photochemical reactions is not immediately available to the complex. We have suggested that the lag may be due to several factors, including the leaving time of the quinol from the reaction center, the diffusion time to the complex, and the time for the head group to cross the membrane. We have suggested aminimal value for the diffusion coefficient of ubiquinone in the membrane (assuming that the lag is due entirely to diffusion) of about 10–9 cm–2 sec–1. The lag is reduced to about 100 µsec when the pool is significantly reduced, showing that quinol from the pool is more rapidly available to the complex than that from the reaction center. With the pool oxidized, similar kinetics are seen when the reduction of cytb 561 occurs through the AA-sensitive site (with reactions at the quinol oxidizing site blocked by myxothiazol). These results show that there is no preferential reaction pathway for transfer of reducing equivalents from reaction center tob/c 1 complex. Oxidation of cytb 561 through the AA-sensitive site can be assayed from the slow phase of the carotenoid electrochromic change, and by comparison with the kinetics of cytb 561. As long as the quinone pool is significantly oxidized, the reaction is not rate-determining for the electrogenic process. On reduction of the pool below 1 quinone per complex, a slowing of the electrogenic process occurs, which could reflect a dependence on the concentration of quinone. If the process is second-order, the rate constant must be about 2–5 times greater than that for quinol oxidation, since the effect on rate is relatively small compared with the effect seen at the quinol oxidizing site when the quinol concentration is changed over theE h range where the first few quinols are produced on reductive titration. When the quinone pool is extracted (experiments in collaboration with G. Venturoli and B. A. Melandri), the slowing of the electrochromic change on reduction of the pool is not enhanced; we assume that this is due to the fact that a minimum of one quinone per active complex is produced by turnover of the quinol oxidizing site. Two lines of research lead us to revise our previous estimate for the minimal value of the quinone diffusion coefficient. These relate to the relation between the diffusion coefficient and the rate constants for processes involving the quinones: (a) The estimated rate constant for reaction of quinone at the AA-site approaches the calculated diffusion limited rate constant, implying an improbably efficient reaction. (b) From a preliminary set of experiments, the activation energy determined by measuring the variation of the rate constant for quinol oxidation with temperature, is about 8 kcal mol–1. Although we do not know the contribution of entropic terms to the pre-exponential factor, the result is consistent with a considerably larger value for the diffusion coefficient than that previously suggested.  相似文献   

11.
Summary The1H NMR signals of the heme methyl, propionate and related chemical groups of cytochromec 3 fromDesulfovibrio vulgaris Miyazaki F (D.v. MF) were site-specifically assigned by means of ID NOE, 2D DQFCOSY and 2D TOCSY spectra. They were consistent with the site-specific assignments of the hemes with the highest and second-lowest redox potentials reported by Fan et al. (Biochemistry,29 (1990) 2257–2263). The site-specific heme assignments were also supported by NOE between the methyl groups of these hemes and the side chain of Val18. All the results contradicted the heme assignments forD.v. MF cytochromec 3 made on the basis of electron spin resonance (Gayda et al. (1987)FEBS Lett.,217 57–61). Based on these assignments, the interaction of cytochromec 3 withD.v. MF ferredoxin I was investigated by NMR. The major interaction site of cytochromec 3 was identified as the heme with the highest redox potential, which is surrounded by the highest density of positive charges. The stoichiometry and association constant were two cytochromec 3 molecules per monomer of ferredoxin I and 108 M–2 (at 53 mM ionic strength and 25°C), respectively.  相似文献   

12.
The temperature dependence of the partial reactions leading to turn-over of the UQH2:cyt c 2 oxidoreductase of Rhodobacter sphaeroides have been studied. The redox properties of the cytochrome components show a weak temperature dependence over the range 280–330 K, with coefficients of about 1 m V per degree; our results suggest that the other components show similar dependencies, so that no significant change in the gradient of standard free-energy between components occurs over this temperature range. The rates of the reactions of the high potential chain (the Rieske iron sulfur center, cytochromes c 1 and c 2, reaction center primary donor) show a weak temperature dependence, indicating an activation energy < 8 kJ per mole for electron transfer in this chain. The oxidation of ubiquinol at the Qz-site of the complex showed a strong temperature dependence, with an activation energy of about 32 kJ mole–1. The electron transfer from cytochrome b-566 to cytochrome b-561 was not rate determining at any temperature, and did not contribute to the energy barrier. The activation energy of 32 kJ mole–1 for quinol oxidation was the same for all states of the quinone pool (fully oxidized, partially reduced, or fully reduced before the flash). We suggest that the activation barrier is in the reaction by which ubiquinol at the catalytic site is oxidized to semiquinone. The most economical scheme for this reaction would have the semiquinone intermediate at the energy level indicated by the activation barrier. We discuss the plausibility of this simple model, and the values for rate constants, stability constant, the redox potentials of the intermediate couples, and the binding constant for the semiquinone, which are pertinent to the mechanism of the ubiquinol oxidizing site.Abbreviations (BChl)2 P870, primary donor of the photochemical reaction center - b/c 1 complex ubiquinol: cytochrome c 2 oxidoreductase - cyt b H cytochrome b-561 or higher potential cytochrome b - cyt b L cytochrome b-566, or low potential cytochrome b - cyt c 1, cyt c 2, cyt c t cytochromes c 1 and c 2, and total cytochrome c (cyt c 1 and cyt c 2) - Fe.S Rieske-type iron sulfur center, Q - QH2 ubiquinone, ubiquinol - Qz, QzH2, Qz ubiquinone, ubiquinol, and semiquinone anion of ubiquinone, bound at quinol oxidizing site - Qz-site ubiquinol oxidizing site (also called Qo-(outside) - Qo (Oxidizing) - QP (Positive proton potential) site) - Qc-site uubiquinone reductase site (also called the Qi-(inside) - QR (Reducing), or - QN (Negative proton potential) site) - UHDBT 5-(n-undecyl)-6-hydroxy-4,7-dioxobenzothiazol  相似文献   

13.
A family of 10 competing, unstructured models has been developed to model cell growth, substrate consumption, and product formation of the pyruvate producing strain Escherichia coli YYC202 ldhA::Kan strain used in fed-batch processes. The strain is completely blocked in its ability to convert pyruvate into acetyl-CoA or acetate (using glucose as the carbon source) resulting in an acetate auxotrophy during growth in glucose minimal medium. Parameter estimation was carried out using data from fed-batch fermentation performed at constant glucose feed rates of qVG=10 mL h–1. Acetate was fed according to the previously developed feeding strategy. While the model identification was realized by least-square fit, the model discrimination was based on the model selection criterion (MSC). The validation of model parameters was performed applying data from two different fed-batch experiments with glucose feed rate qVG=20 and 30 mL h–1, respectively. Consequently, the most suitable model was identified that reflected the pyruvate and biomass curves adequately by considering a pyruvate inhibited growth (Jerusalimsky approach) and pyruvate inhibited product formation (described by modified Luedeking–Piret/Levenspiel term).List of symbols cA acetate concentration (g L–1) - cA,0 acetate concentration in the feed (g L–1) - cG glucose concentration (g L–1) - cG,0 glucose concentration in the feed (g L–1) - cP pyruvate concentration (g L–1) - cP,max critical pyruvate concentration above which reaction cannot proceed (g L–1) - cX biomass concentration (g L–1) - KI inhibition constant for pyruvate production (g L–1) - KIA inhibition constant for biomass growth on acetate (g L–1) - KP saturation constant for pyruvate production (g L–1) - KP inhibition constant of Jerusalimsky (g L–1) - KSA Monod growth constant for acetate (g L–1) - KSG Monod growth constant for glucose (g L–1) - mA maintenance coefficient for growth on acetate (g g–1 h–1) - mG maintenance coefficient for growth on glucose (g g–1 h–1) - n constant of extended Monod kinetics (Levenspiel) (–) - qV volumetric flow rate (L h–1) - qVA volumetric flow rate of acetate (L h–1) - qVG volumetric flow rate of glucose (L h–1) - rA specific rate of acetate consumption (g g–1 h–1) - rG specific rate of glucose consumption (g g–1 h–1) - rP specific rate of pyruvate production (g g–1 h–1) - rP,max maximum specific rate of pyruvate production (g g–1 h–1) - t time (h) - V reaction (broth) volume (L) - YP/G yield coefficient pyruvate from glucose (g g–1) - YX/A yield coefficient biomass from acetate (g g–1) - YX/A,max maximum yield coefficient biomass from acetate (g g–1) - YX/G yield coefficient biomass from glucose (g g–1) - YX/G,max maximum yield coefficient biomass from glucose (g g–1) - growth associated product formation coefficient (g g–1) - non-growth associated product formation coefficient (g g–1 h–1) - specific growth rate (h–1) - max maximum specific growth rate (h–1)  相似文献   

14.
(1) The electron transport system of heterotrophically dark-grown Rhodobacter capsulatus was investigated using the wild-type strain MT1131 and the phototrophic non-competent (Ps-) mutant MT-GS18 carrying deletions of the genes for cytochrome c 1 and b of the bc 1 complex and for cytochrome c 2. (2) Spectroscopic and thermodynamic data demonstrate that deletion of both bc 1 complex and cyt. c 2 still leaves several haems of c- and b-type with Em7.0 of +265 mV and +354 mV at 551–542 nm, and +415 mV and +275 mV at 561–575 nm, respectively. (3) Analysis of the oxidoreduction kinetic patterns of cytochromes indicated that cyt. b 415 and cyt. b 275 are reduced by either ascorbate-diaminodurene or NADH, respectively. (4) Growth on different carbon and nitrogen sources revealed that the membrane-bound electron transport chain of both MT1131 and MT-GS18 strains undergoes functional modifications in response to the composition of the growth medium used. (5) Excitation of membrane fragments from cells grown in malate minimal medium by a train of single turnover flashes of light led to a rapid oxidation of 32% of the membrane-bound c-type haem complement. Conversely, membranes prepared from peptone/yeast extract grown cells did not show cyt. c photooxidation. These results are discussed within the framework of an electron transport chain in which alternative pathways bypassing both the cyt. c 2 and bc 1 complex might involve high-potential membrane bound haems of b- and c-type.Abbreviations AA antimycin A - CCCP carbonylcyanide m-chlorophenyl hydrazone - CN- cyanide - DAD diaminodurene - Q2H2 ubiquinol-2 - Q-pool ubiquinone-10 pool - RC photochemical reaction center  相似文献   

15.
Calcium binding to spinach (Spinacia oleracea L.) stromal proteins was examined by dual-wavelength spectrophotometry using the metallochromic indicator tetramethylmurexide. The data are consistent with the existence of at least two, probably independent, classes of binding sites. The total number of binding sites varied between 90–155 nmol·mg–1 protein with average binding constants of 1.1–2.7·mM–1. Both Mg2+ and La3+ inhibited calcium binding competitively, with average inhibitor constants of 0.26·mM–1 and 39.4·mM–1, respectively; an increase in the potassium concentration up to 50 mM had no effect. In a typical experiment a decrease in pH (7.8 to 7.1) resulted in a decrease in the total number of calcium binding sites from 90 to 59 nmol·mg–1 protein, but in an increase of the average affinity from 2.7 to 4.5·mM–1. Calculations, using these data and those of Gross and Hess (1974, Biochim. Biophys. Acta 339, 334–346) for binding site I of washed thylakoid membranes, showed that the free-Ca2+ concentration in the stroma under dark conditions, pH 7.1, is higher than under light conditions, pH 7.8. The physiological relevance of the observed calcium binding by stromal proteins is discussed.Abbreviations Ca b 2+ bound calcium - Ca f 2+ free calcium  相似文献   

16.
The response of stomata in isolated epidermis to the concentration of CO2 in the gaseous phase was examined in a C3 species, the Argenteum mutant of Pisum sativum, and a crassulacean-acid-metabolism (CAM) species, Kalanchoë daigremontiana. Epidermis from leaves of both species was incubated on buffer solutions in the presence of air containing various volume fractions of CO2 (0 to 10000·10–6). In both species and in the light and in darkness, the effect of CO2 was to inhibit stomatal opening, the maximum inhibition of opening occurring in the range 0 to 360·10–6. The inhibition of opening per unit change in concentration was greatest between volume fractions of 0 and 240·10–6. There was little further closure above the volume fraction of 360·10–6, i.e. approximately ambient concentration of CO2. Thus, although leaves of CAM species may experience much higher internal concentrations of CO2 in the light than those of C3 plants, this does not affect the sensitivity of their stomata to CO2 concentration or the range over which they respond. Stomatal responses to CO2 were similar in both the light and the dark, indicating that effects of CO2 on stomata occur via mechanisms which are independent of light. The responses of stomata to CO2 in the gaseous phase took place without the treatments changing the pH of the buffered solutions. Thus it is unlikely that CO2 elicited stomatal movement by changing either the pH or the HCO 3 /CO 3 2- equilibria. It is suggested that the concentration of dissolved unhydrated CO2 may be the effector of stomatal movement and that its activity is related to its reactivity with amines.  相似文献   

17.
The midpoint potential of the [2Fe–2S] cluster of the Rieske iron–sulfurprotein (E m 7 = +280mV) is the primary determinant of the rate of electron transfer from ubiquinol to cytochromec catalyzed by the cytochrome bc 1 complex. As the midpoint potential of the Rieske clusteris lowered by altering the electronic environment surrounding the cluster, theubiquinol-cytochrome c reductase activity of the bc 1 complex decreases; between 220 and 280 mV therate changes 2.5-fold. The midpoint potential of the Rieske cluster also affects thepresteady-state kinetics of cytochrome b and c 1 reduction. When the midpoint potential of the Rieskecluster is more positive than that of the heme of cytochrome c 1, reduction of cytochrome bis biphasic. The fast phase of b reduction is linked to the optically invisible reduction of theRieske center, while the rate of the second, slow phase matches that of c 1 reduction. The ratesof b and c 1 reduction become slower as the potential of the Rieske cluster decreases andchange from biphasic to monophasic as the Rieske potential approaches that of theubiquinone/ubiquinol couple. Reduction of b and c 1 remain kinetically linked as the midpoint potentialof the Rieske cluster is varied by 180 mV and under conditions where the presteady statereduction is biphasic or monophasic. The persistent linkage of the rates of b and c 1 reduction isaccounted for by the bifurcated oxidation of ubiquinol that is unique to the Q-cycle mechanism.  相似文献   

18.
The mitochondrial cytochrome bc 1 complex is a multifunctional membrane protein complex. Itcatalyzes electron transfer, proton translocation, peptide processing, and superoxide generation.Crystal structure data at 2.9 Å resolution not only establishes the location of the redox centersand inhibitor binding sites, but also suggests a movement of the head domain of the iron–sulfurprotein (ISP) during bc 1 catalysis and inhibition of peptide-processing activity during complexmaturation. The functional importance of the movement of extramembrane (head) domain ofISP in the bc 1 complex is confirmed by analysis of the Rhodobacter sphaeroides bc 1 complexmutants with increased rigidity in the ISP neck and by the determination of rate constants foracid/base-induced intramolecular electron transfer between [2Fe–2S] and heme c 1 in nativeand inhibitor-loaded beef complexes. The peptide-processing activity is activated in bovineheart mitochondrial bc 1 complex by nonionic detergent at concentrations that inactivate electrontransfer activity. This peptide-processing activity is shown to be associated with subunits Iand II by cloning, overexpression and in vitro reconstitution. The superoxide-generation siteof the cytochrome bc 1 complex is located at reduced b L and Q. The reaction is membranepotential-, and cytochrome c-dependent.  相似文献   

19.
The ethylene-binding site (EBS) from Phaseolus vulgaris cv. Canadian Wonder cotyledons can be solubilised from 96,000 g pelleted material by Triton X-100 or sodium cholate. Extraction of 96,000 g pellets with acetone, butanol or butanol and ether results in a total loss of ethylene-binding activity. Like the membrane-bound form, the solubilised EBS has an apparent KD(liquid) of 10-10 M at a concentration of 32 pmol EBS per gram tissue fresh weight. Propylene and acetylene act as competitive inhibitors, carbon dioxide appears to promote ethylene binding and ethane has no significant effect. The solubilised EBS is completely denatured affect. The solubilised EBS is completely denatured after 10 min at 70°C, by 1 mM mercaptoethanol and 0.1 mM dithiothreitol, but not by trypsin or chymotrypsin. However, solubilisation decreases the rate constant of association from 103 M-1 s-1 to 101–102 M-1 s-1 and hence does not permit experimental determination of the rate constant of dissociation. The pH optimum for ethylene binding is altered from the range pH 7–10 in the membrane-bound form to the pH range 4–7 in the solubilised form. The EBS appears to be a hydrophobic, intergral membrane protein, which requires a hydrophobic environment to retain its activity. Partitioning of the EBS into polymer phases is determined by the detergent used for solubilisation indicating that when solubilised, the EBS forms a complex with detergent molecules.Abbreviations EBS ethylene-binding site - PEG polyethylene glycol  相似文献   

20.
Cytochrome c (cyt c) was reduced by a tyrosine-containing peptide, tyrosyltyrosylphenylalanine (TyrTyrPhe), at pH 6.0–8.0, while tyrosinol or tyrosyltyrosine (TyrTyr) could not reduce cyt c effectively under the same condition. Cyt c was reduced at high peptide concentration, whereas the reaction did not occur effectively at low concentration. The reaction rate varied with time owing to a decrease in the TyrTyrPhe concentration and the production of tyrosine derivatives during the reaction. The initial rate constants were 2.4×10–4 and 8.1×10–4 s–1 at pH 7.0 and 8.0, respectively, for the reaction with 1.0 mM TyrTyrPhe in 10 mM phosphate buffer at 15°C. The reciprocal initial rate constant (1/kint) increased linearly against the reciprocal peptide concentration and against the linear proton concentration, whereas logkint decreased linearly against the root of the ionic strength. These results show that deprotonated (TyrTyrPhe), presumably deprotonated at a tyrosine site, reduces cyt c by formation of an electrostatic complex. No significant difference in the reaction rate was observed between the reaction under nitrogen and oxygen atmospheres. From the matrix-assisted laser desorption ionization time-of-flight mass spectra of the reaction products, formation of a quinone and other tyrosine derivatives of the peptide was supported. These products should have been produced from a tyrosyl radical. We interpret the results that a cyt cox/(TyrTyrPhe)cyt cred/(TyrTyrPhe) equilibrium is formed, which is usually shifted to the left. This equilibrium may shift to the right by reaction of the produced tyrosyl radical with the tyrosine sites of unreacted TyrTyrPhe peptides.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号